Optical Nonlinearities in Chalcogenide Glasses and their Applications (Springer Series in Optical Sciences)

  • 22 47 4
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up

Optical Nonlinearities in Chalcogenide Glasses and their Applications (Springer Series in Optical Sciences)

Springer Series in optical sciences founded by H.K.V. Lotsch Editor-in-Chief: W. T. Rhodes, Atlanta Editorial Board: A.

593 34 3MB

Pages 210 Page size 400 x 619 pts Year 2011

Report DMCA / Copyright

DOWNLOAD FILE

Recommend Papers

File loading please wait...
Citation preview

Springer Series in

optical sciences founded by H.K.V. Lotsch Editor-in-Chief: W. T. Rhodes, Atlanta Editorial Board: A. Adibi, Atlanta T. Asakura, Sapporo T. W. H¨ansch, Garching T. Kamiya, Tokyo F. Krausz, Garching B. Monemar, Link¨oping M. Ohtsu, Tokyo H. Venghaus, Berlin H. Weber, Berlin H. Weinfurter, M¨unchen

135

Springer Series in

optical sciences The Springer Series in Optical Sciences, under the leadership of Editor-in-Chief William T. Rhodes, Georgia Institute of Technology, USA, provides an expanding selection of research monographs in all major areas of optics: lasers and quantum optics, ultrafast phenomena, optical spectroscopy techniques, optoelectronics, quantum information, information optics, applied laser technology, industrial applications, and other topics of contemporary interest. With this broad coverage of topics, the series is of use to all research scientists and engineers who need up-to-date reference books. The editors encourage prospective authors to correspond with them in advance of submitting a manuscript. Submission of manuscripts should be made to the Editor-in-Chief or one of the Editors. See also www.springeronline.com/series/624

Editor-in-Chief William T. Rhodes Georgia Institute of Technology School of Electrical and Computer Engineering Atlanta, GA 30332-0250, USA E-mail: [email protected]

Editorial Board Ali Adibi Georgia Institute of Technology School of Electrical and Computer Engineering Atlanta, GA 30332-0250, USA E-mail: [email protected]

Toshimitsu Asakura Hokkai-Gakuen University Faculty of Engineering 1-1, Minami-26, Nishi 11, Chuo-ku Sapporo, Hokkaido 064-0926, Japan E-mail: [email protected]

Theodor W. H¨ansch Max-Planck-Institut f¨ur Quantenoptik Hans-Kopfermann-Straße 1 85748 Garching, Germany E-mail: [email protected]

Takeshi Kamiya Ministry of Education, Culture, Sports Science and Technology National Institution for Academic Degrees 3-29-1 Otsuka, Bunkyo-ku Tokyo 112-0012, Japan E-mail: [email protected]

Ferenc Krausz Ludwig-Maximilians-Universit¨at M¨unchen Lehrstuhl f¨ur Experimentelle Physik Am Coulombwall 1 85748 Garching, Germany and Max-Planck-Institut f¨ur Quantenoptik Hans-Kopfermann-Straße 1 85748 Garching, Germany E-mail: [email protected]

Bo Monemar Department of Physics and Measurement Technology Materials Science Division Link¨oping University 58183 Link¨oping, Sweden E-mail: [email protected]

Motoichi Ohtsu University of Tokyo Department of Electronic Engineering 7-3-1 Hongo, Bunkyo-ku Tokyo 113-8959, Japan E-mail: [email protected]

Herbert Venghaus Fraunhofer Institut f¨ur Nachrichtentechnik Heinrich-Hertz-Institut Einsteinufer 37 10587 Berlin, Germany E-mail: [email protected]

Horst Weber Technische Universit¨at Berlin Optisches Institut Straße des 17. Juni 135 10623 Berlin, Germany E-mail: [email protected]

Harald Weinfurter Ludwig-Maximilians-Universit¨at M¨unchen Sektion Physik Schellingstraße 4/III 80799 M¨unchen, Germany E-mail: [email protected]

A. Zakery

S.R. Elliott

Optical Nonlinearities in Chalcogenide Glasses and their Applications With 92 Figures, 14 in Color and 1 8 Tables

123

Dr. A. Zakery Shiraz University, College of Sciences, Department of Physics Golestan Avenue, Shiraz 71454, Iran E-mail: [email protected]

Professor Dr. S.R. Elliott University of Cambridge, Department of Chemistry Lensfield Road, Cambridge CB2 1EW, UK mail: [email protected]

ISSN 0342-4111 ISBN-10 3-540-71066-3 Springer Berlin Heidelberg New York ISBN-13 978-3-540-71066-0 Springer Berlin Heidelberg New York Library of Congress Control Number:

2007923601

This work is subject to copyright. All rights are reserved, whether the whole or part of the material is concerned, specif ically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microf ilm or in any other way, and storage in data banks. Duplication of this publication or parts thereof is permitted only under the provisions of the German Copyright Law of September 9, 1965, in its current version, and permission for use must always be obtained from Springer-Verlag. Violations are liable to prosecution under the German Copyright Law. Springer is a part of Springer Science+Business Media. springer.com © Springer-Verlag Berlin Heidelberg 2007 The use of general descriptive names, registered names, trademarks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. Typesetting by SPi using Springer LATEX macro package Cover: eStudio Calamar Steinen Printed on acid-free paper

SPIN: 11495802

56/3180/SPI 5 4 3 2 1 0

Preface

Photonics, which uses photons for information and image processing, is labeled as one of the technologies of the 21st century, for which nonlinear optical processes provide the key functions of frequency conversion and optical switching. Chalcogenide glasses are based on the chalcogen elements S, Se, and Te. These glasses are formed by the addition of other elements such as Ge, As, Sb, Ga, etc. These glasses are low-phonon energy materials and are generally transparent from the visible to infrared. Chalcogenide glasses can be doped by rare-earth elements such as Er, Nd, Pr, etc., and hence numerous applications of active optical devices have been proposed. These glasses are optically highly nonlinear and could therefore be useful for all-optical switching. This book is a review of recent progress in the science and technology of chalcogenide glasses, with an emphasis on their nonlinear optical properties, for graduate students, practising engineers and scientists from a wide multidisciplinary area such as physics, chemistry, electrical engineering and material science. Since the interest in this area is growing worldwide, a book dealing with this subject will be of great value to researchers of varied backgrounds. Chalcogenide glasses and their electronic, structural, and photoinduced properties are introduced. Techniques to characterize the linear and nonlinear optical properties of these glasses are introduced and used to measure the optical constants of chalcogenide glasses in the form of bulk, thin film and fiber. The possibilities of fabricating passive and active devices are presented. A novel application of chalcogenide glasses, namely all-optical switching for the fabrication of efficient femtosecond switches, is introduced. Finally other applications of chalcogenide glasses, such as optical limiting, second-harmonic generation, fabrication of rib and ridge waveguides and of fiber gratings, optical regenerators and the possibility of using these glasses in all-optical nonlinear integrated circuits and the possibility of enhancing optical nonlinearities by inclusion of nanometals, are discussed in some detail. We wish to express gratitude to our families and especially to our wives Susan and Penny who, in spite of their own professional schedules, have

VI

Preface

provided valuable support and understanding for this project. We would also like to thank many of our colleagues and students from whom we benefited very much from their collaboration. A.Z. would also like to thank the Royal Society of London for providing support for a short visit to Cambridge which helped to collect part of the literature materials in an early phase of the writing of this book.

Shiraz and Cambridge April 2007

A. Zakery S.R. Elliott

Contents

1

An 1.1 1.2 1.3

1.4 1.5 1.6 1.7 1.8

1.9 2

Introduction to Chalcogenide Glasses . . . . . . . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Structure of Chalcogenide Glasses . . . . . . . . . . . . . . . . . . . . . . . . . Electronic Properties of Chalcogenide Glasses . . . . . . . . . . . . . . . 1.3.1 Electronic States in Chalcogenide Glasses . . . . . . . . . . . . 1.3.2 Measurements of the Absorption Coefficient and the Optical Gap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chalcogenide Glasses for Near-Infrared Optics . . . . . . . . . . . . . . Chalcogenide Glasses for Mid-IR and Far-IR Applications . . . . Bulk Chalcogenide Glasses, Composition, and Optical Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chalcogenide Thin Films and Comparison with the Bulk . . . . . Photoinduced Changes in Chalcogenide Glasses . . . . . . . . . . . . . 1.8.1 Photoinduced Phenomena . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.2 Exposure Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.3 Measurements of the Propagation Losses by a Prism Coupler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.4 Measurements of Propagation Losses in Laser-Written Waveguides . . . . . . . . . . . . . . . . . . . . . . . . Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Basic Concepts of Nonlinear Optics . . . . . . . . . . . . . . . . . . . . . . . 2.1 Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Wave Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Linear Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.2 Nonlinear Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 The Harmonic Oscillator Model in Linear Optics . . . . . . . . . . . . 2.4 The Anharmonic Oscillator Model in Nonlinear Optics . . . . . . . 2.5 Properties of Anisotropic Media . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6 Second-Harmonic Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1 1 1 6 6 8 10 12 14 17 21 21 23 25 26 27 29 29 30 31 34 37 40 42 43

VIII

3

Contents

2.7 Self-Phase Modulation and Soliton Generation . . . . . . . . . . . . . . 2.7.1 Optical Solitons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.7.2 Mechanisms of Nonlinearity . . . . . . . . . . . . . . . . . . . . . . . . 2.7.3 Optical Phase Conjugation . . . . . . . . . . . . . . . . . . . . . . . . . 2.7.4 Optical Bistability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.7.5 Stimulated Raman Scattering . . . . . . . . . . . . . . . . . . . . . . . 2.7.6 Third-Harmonic Generation . . . . . . . . . . . . . . . . . . . . . . . .

44 45 47 48 50 51 52

Experimental Techniques to Measure Nonlinear Optical Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Degenerate Four-Wave Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Nearly Degenerate Three-wave Mixing . . . . . . . . . . . . . . . . . . . . . 3.4 Z-Scan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5 Third-Harmonic Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6 Optical Kerr Gate and Ellipse Rotation . . . . . . . . . . . . . . . . . . . . 3.6.1 Optical Kerr Gate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6.2 Ellipse Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7 Self-Phase Modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.8 Spectrally Resolved Two-Beam Coupling . . . . . . . . . . . . . . . . . . 3.9 Mach-Zehnder Interferometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

55 55 55 59 61 63 64 64 66 67 69 70 73

4

Measurement of Nonlinear Optical Constants . . . . . . . . . . . . . 75 4.1 Measurements of Nonlinear Refractive Index n2 . . . . . . . . . . . . . 75 4.2 Measurements of Nonlinear Absorption Coefficient β . . . . . . . . 91 4.3 Determination of Three Photon-Absorption and Multiphoton Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 4.4 Second-Harmonic Generation, Phase Conjugation, etc . . . . . . . . 95 4.5 Comparison of Chalcogenide Nonlinearities with Silica . . . . . . . 102

5

Optical Nonlinearities in Chalcogenide Fibres . . . . . . . . . . . . . . 107 5.1 Fabrication of Chalcogenide Fibers and Their Linear Optical Properties . . . . . . . . . . . . . . . . . . . . . . . 107 5.1.1 Fabrication of Fibers by Extrusion . . . . . . . . . . . . . . . . . . 108 5.1.2 Physical and Linear Optical Properties of Chalcogenide Fibers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 5.2 Nonlinear Optical Properties of Fibers . . . . . . . . . . . . . . . . . . . . . 111 5.2.1 Features of Chalcogenide Glass as a Nonlinear Material . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 5.2.2 Stimulated Light Scattering and Super-Continuum Generation . . . . . . . . . . . . . . . . . . . 112 5.2.3 Second-Order Nonlinearity in Poled Glass . . . . . . . . . . . . 113

Contents

IX

5.3 Pulse Propagation in Fibers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 5.3.1 Propagation of Optical Fields . . . . . . . . . . . . . . . . . . . . . . 114 5.3.2 Nonlinear Pulse Propagation . . . . . . . . . . . . . . . . . . . . . . . . 116 5.3.3 Higher-Order Nonlinear Effects . . . . . . . . . . . . . . . . . . . . . 120 5.4 Group-Velocity Dispersion Compensation by Fiber Gratings . . 121 5.5 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 6

Optical Switching in Chalcogenide Glasses . . . . . . . . . . . . . . . . 129 6.1 Criteria of Material Properties for All-optical Switching . . . . . . 129 6.2 Design Issues for All-Optical Switching . . . . . . . . . . . . . . . . . . . . 131 6.3 All-Optical Switching in Chalcogenide Glasses . . . . . . . . . . . . . . 131 6.3.1 All-Optical Switching using Chalcogenide Glass Fibers . . . . . . . . . . . . . . . . . . . . . 131 6.3.2 All-Optical Switching in Thin Chalcogenide Films . . . . 137 6.4 All-Optical Switches, AND Gate, NOR Gate, etc. . . . . . . . . . . . 145 6.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 6.4.2 Nonlinear Interferometric Devices . . . . . . . . . . . . . . . . . . . 147 6.4.3 Nonlinear Beam-Coupling Devices . . . . . . . . . . . . . . . . . . 147 6.4.4 Polarization Switching Devices . . . . . . . . . . . . . . . . . . . . . 148 6.4.5 Soliton Switching Devices . . . . . . . . . . . . . . . . . . . . . . . . . . 148 6.5 Limitations of All-Optical Switches . . . . . . . . . . . . . . . . . . . . . . . 149 6.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

7

Issues and Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 7.1 Optical Limiting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 7.2 Second-Harmonic Generation and Electro-Optic Effects . . . . . . 153 7.3 Fabrication of Rib and Ridge Waveguides and of Fiber Gratings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 7.4 All-Optical Nonlinear Integrated Circuits . . . . . . . . . . . . . . . . . . 166 7.5 Inclusion of Metal Nanoparticles to Enhance Nonlinearity . . . . 168 7.6 Other Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 7.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195

1 An Introduction to Chalcogenide Glasses

1.1 Introduction Chalcogenide glasses are based on the chalcogen elements S, Se, and Te. These glasses are formed by the addition of other elements such as Ge, As, Sb, Ga, etc. They are low-phonon-energy materials and are generally transparent from the visible up to the infrared. Chalcogenide glasses can be doped by rareearth elements, such as Er, Nd, Pr, etc., and hence numerous applications of active optical devices have been proposed. Since chalcogenide-glass fibers transmit in the IR, there are numerous potential applications in the civil, medical, and military areas. Passive applications utilize chalcogenide fibers as a light conduit from one location to another point without changing the optical properties, other than those due to scattering, absorption, and reflection. These glasses are optically highly nonlinear and could therefore be useful for all-optical switching (AOS). Chalcogenide glasses are sensitive to the absorption of electromagnetic radiation and show a variety of photoinduced effects as a result of illumination. Various models have been put forward to explain these effects, which can be used to fabricate diffractive, waveguide and fiber structures. For recent reviews, see [1–4]. Next-generation devices for telecommunication and related applications will rely on the development of materials which possess optimized physical properties that are compatible with packaging requirements for systems in planar or fiber form. This allows suitable integration to existing fiber-based applications, and hence requires appropriate consideration as to material choice, stability, and long-term aging behavior.

1.2 Structure of Chalcogenide Glasses Solids are a particular state of condensed matter characterized by strong interactions between the constituent particles (atoms, molecules). Solids can be found or prepared either in an ordered (crystalline) state or in a disordered (noncrystalline) state. While the ordered state of a solid is limited to only

2

1 An Introduction to Chalcogenide Glasses

a few structural forms, a disordered material is neither unique nor clearly defined. An ideal crystal corresponds to a regular arrangement of atoms in a lattice with well-defined symmetry, and a structural unit called the unit cell can be defined. Translation of the unit cell along the three coordinate axes reproduces the whole assembly of atoms. A real crystal does not exhibit perfect periodicity in space and contains various kinds of imperfections or defects. Solids which lack the periodicity of the atoms are called noncrystalline solids or amorphous, vitreous or glassy solids. While crystals possess long-range order (LRO), in amorphous materials short-range order (SRO) still exists. Although the first and second nearest-neighbor coordination shells are welldefined, atoms in the third coordination sphere start to become uncorrelated with those in the first one. In other words, the limit of short- and mediumrange order is the first 3–4 interatomic distances. The price to be paid for the loss of LRO is the appearance of fluctuations in angles and distances between the bonds. The ideal noncrystalline network is difficult to define. Particularly, different thermal treatments lead to various noncrystalline arrangements of atoms. A continuous random network [5] might be considered to be an ideal noncrystalline network for covalent solids. The structure of chalcogenide glasses, however, cannot be described by means of a continuous random network which is isotropic in three dimensions, as in the case of amorphous silicon for example. As2 S3 , As2 Se3 , GeS2 , and GeSe2 can be locally layer-like, while pure S and Se are chain like. For all these materials, there is considerable flexibility of the structure as a result of the weak van der Waal’s bonding between layers or chains [6], so that changes in the structure can be relatively easily accommodated. Raman (or inelastic) scattering of light in a material yields structural and dynamic information on a molecular level. The nondestructive nature of the probe, and flexibility in sampling arrangements, has opened up many potentially new areas where Raman measurements have proven valuable [7,8]. Nearinfrared (NIR) Raman spectroscopy can be used in the analysis of materials which are strongly absorbing in the visible. A distinct advantage over the more conventional approach using the visible part of the spectrum as the excitation wavelength is the ability to obtain the Raman spectrum of photosensitive compounds without interference from photoreactions caused by the probe beam. In As–S–Se chalcogenide glasses, shifting the excitation wavelength to 840 nm (typically below the band gap) allows one to obtain high-quality Raman spectra without material modification. Raman spectroscopy can be extremely powerful in the microstructural analysis of single and multilayer waveguide devices [9]. Here, the material of interest is made in the form of a slab waveguide, thereby significantly increasing both the scattering volume and the electric-field intensity within the film. Raman scattering in the NIR can be excited with 840 nm radiation from a tunable Ti:sapphire laser (30– 50 mW). The dominant feature in binary sulfide and selenide compounds are bands at 345 cm−1 (As40 S60 ) and 230 cm−1 (As40 Se60 ), respectively, [10] (see Fig. 1.1).

1.2 Structure of Chalcogenide Glasses

Raman Intensity (arb. units)

As40Se60

lexcitation = 840 nm

a b

As40S30Se30

c

As40S45Se15

d

As40S60

e-f

As32S34Se34

 10

As24S38Se38

100

3

200

300

400

500

−1

Frequency (cm ) Fig. 1.1. Raman spectra of bulk glasses obtained with near-infrared excitation (λ = 840 nm) as a function of compositional variation. Spectral resolution is 1.5 cm−1 (after [10])

A strong broad band is seen which is attributed to the antisymmetric As–(S,Se)–As stretching vibration in As(S, Se)3 pyramidal units. In ternary compounds with a S/Se = 1 molecular ratio and decreasing As content, a progressive decrease in the intensity of these broad bands is observed, indicative of a decrease in the number of As-containing pyramidal sites. New bands appearing around 255 cm−1 and 440–480 cm−1 form in chalcogen-rich glasses, and are attributed to Se–Se homopolar bonds. These units serve as chalcogen chains connecting the remaining pyramidal units. The small numbers of S–S bonds, indicated by a weak band near 495 cm−1 , for equal concentrations of S and Se suggests that S stays with the remaining pyramids and that it is the Se which dominates the connecting chain units. Deviations from bulk glass properties in fibers and films are observed. The extent of such compositional variation, and the resulting structural units formed, varies with the specific fiber or film-processing technique used. Neutron-scattering and X-ray diffraction studies on bulk sulfide and selenide glasses and their thin films [11] have shown variations in structural units on the intermediate-range order scale. Depending on processing conditions, polymeric cages (based on As4 S4 units) or less connected groups of As–S pyramidal units, were observed. Furthermore, these units are much more metastable and can be structurally modified or eliminated with postdeposition processing. Waveguide Raman spectroscopy (WRS) has been applied to the structural characterization of chalcogenide

1 An Introduction to Chalcogenide Glasses

Raman Intensity (arb. units)

4

Waveguide Raman Spectra λ = 840 nm

Annealed As-deposited

Photoinduced 200

250

300

350

400

450

500

Frequency (cm−1) Fig. 1.2. Variation in waveguide Raman spectra for a fresh, annealed, and photostructurally modified As2 S3 channel waveguide (exposure λ = 514.5 nm). Excitation wavelength was 840 nm (after [10])

glasses [12, 13]. The excitation beam (840 nm) was launched into the end face of an As2 S3 channel waveguide at various lateral positions. Although the As2 S3 film had a thickness of just 1.5 µm, the high signal-to-noise ratio achieved by guided-mode excitation was evident and low-frequency Raman peaks were well separated from interfering Rayleigh scattering. In comparison with the bulk, new substructures appeared in the film spectra as compared with the broad features of the bulk spectra (see Fig. 1.2). These differences in the spectra from the bulk are due to different (molecular) arrangements of the constituent atoms within the films. These sharp, molecular-signature features were confirmed not to be due to crystallinity within the film, but most likely result from the formation of as-deposited As4 S4 units [11, 14]. Rutherford backscattering spectroscopy (RBS) is an analytical tool that gives very useful information regarding compositional and structural analysis of films, as well as a precise measurement of the film thickness. Results obtained showed no apparent variation in composition and small (less than 10%) density variation in single-layer As2 S3 films [15]. Multilayer films, whose thickness can be measured using SEM images, display compositional and density modifications associated with the annealing process. Film ageing was investigated in films after almost a year. Stoichiometric and thickness modifications, caused by ageing, were observed in unannealed structures [15]. No apparent changes were detected in annealed films. The RBS data shows that the ratio of the sulfur-to-arsenic concentration increases during the annealing process. This suggests that the sulfur is not evaporating during annealing. However, SEM shows a modification in the layer thickness for the multilayer structure. This result implies that the molecules in the films are rearranging. NIR Raman results complement this conclusion.

1.2 Structure of Chalcogenide Glasses

5

Chalcogenide glasses seem to experience slight modifications with time, under standard conditions (i.e. room temperature). However, these structural modifications or relaxation in the molecular structure of the films are less apparent in annealed structures, which suggests that annealed films are fairly stable in time. Seal et al. [16] have used X-ray photoelectron spectroscopy (XPS) to study the resulting chemical composition of As2 S3 at the film surface as compared to the parent bulk glass and the corresponding variation in the nature of chemical bonds and electronic structure. Ar+ -ion sputtering converts the amorphous phase into a crystalline phase as the binding energy of the As peak increases from 42.7 to 43.8 eV [16]. Inference of the crystalline phase in the sputtered region is based on the fact that 99.99% pure melted As2 S3 glass has a similar binding energy to that seen for As in its crystalline form. It was found that laser irradiation induces structural and chemical changes in the sample, as can be seen from the change in the As/S ratio with illumination. Polarization direction, vertical and circular polarization were found to induce an As-deficient structure and horizontal polarization was found to make the system As-rich [16]. It was also found that, when films are illuminated, some non-bridging S atoms were observed but no nonbridging atoms appeared to be created when the sample was sputtered with an Ar+ beam [16]. EXAFS studies [17] have shown that there is chemical disorder in the structural network of GaLaS thin films, although chemical ordering is predominant in bulk GaLaS glass. EXAFS results show that gallium is always fourfold coordinated in the GLS network. Both Ga–Ga and S–S bonds occur in GLS thin-film samples but the lanthanum atoms remain eightfold coordinated by sulfur alone. The bond lengths were the same as those found in bulk glass. Although the nearest-neighbor environment is well-defined, there is considerable bond-angle variation, and hence a wide variation in second-neighbor distances. Benazeth et al. [18] also studied the structure of bulk GLS glasses using EXAFS at the gallium K edge and lanthanum L3 edge. Their results show that gallium atoms in the glass exist in tetrahedral networks of GaS4 , and the Ga–S distances in the glasses are identical to those existing in the crystalline form of Ga2 S3 . The gallium and sulfur environment in crystalline Ga2 S3 is such that two of the three sulfur atoms are linked to three gallium atoms and the third sulfur atom is linked to two gallium atoms. The bonds that link two of the three sulfur atoms to three gallium atoms consist of two covalent bonds and a third dative bond, while the third sulfur atom linked to two gallium atoms represents the bridging atom. The addition of La2 S3 brings in an additional S 2− anion that results in modification of the dative bond of the trigonally coordinated sulfur atom. The dative bond is broken and the S 2− anion provided by the modifying rare-earth sulfide helps in restoring and maintaining the tetrahedral environment of GaS4 , at the same time creating a negative site for the La3+ cation. Lucazeau et al. [19] studied the structure of these glasses using Raman spectroscopy. The Raman spectra of the glasses and of similar crystalline phases were compared and spectral differences were found between the two glassy and crystalline states. This has been interpreted in

6

1 An Introduction to Chalcogenide Glasses

terms of structural modification of the short-range periodicity around the Ga atoms, although there was no conclusive evidence.

1.3 Electronic Properties of Chalcogenide Glasses Chalcogenide glasses can be characterized as being variously covalent, metallic, and ionic. In covalent chalcogenide glasses such as Se and As2 S3 , the so-called 8-N rule applies to the coordination number of the constituent atoms, e.g. the coordination number of chalcogens is generally 2 since the total number of valence electrons is N = 6. The magnitude of the band gap is 1–3 eV depending on the composition, the band gap increasing in the series Te → Se → S. Electrical conduction in many chalcogenide glasses is governed by holes. Accordingly, these glasses can be regarded as amorphous semiconductors. However, in a glass containing large amounts of Te, the band gap decreases (∼1 eV), and the metallic character increases. Moreover, in glasses such as Ag–As(Ge)–S, the coordination number of S is demonstrated to be 3–4 [20], and ionic conduction of Ag+ governs the electrical conductivity. So these glasses can be considered as ionic glasses or ion-conducting amorphous semiconductors. 1.3.1 Electronic States in Chalcogenide Glasses If we compare the optical properties of crystalline and amorphous As2 S3 and As2 Se3 , we see that the effect of disorder on the electronic structure is relatively small; the optical absorption edges (band gaps) are very similar. The major difference between crystalline and amorphous solids lies in a higher density of traps and a larger energy distribution of the trapping levels in the amorphous solid. The electrical band gap for As2 Se3 is 1.1 eV in comparison to 1.9 eV for the optical gap. The optical gap is therefore approximately double the electrical gap, which indicates that the Fermi level is situated near the middle of the mobility gap (the energy interval between the demarcation levels between localized and extended states in valence and conduction bands). In a chalcogen, e.g. selenium, the four 4p electrons occupy two bonding orbitals representing covalent bonding and one orbital named a lone-pair orbital, which does not participate in the covalent bonding. The p electrons give rise to strong covalent bonds. Lone-pair electrons determine the dihedral angles and, being the uppermost electrons in the valence band, play an important role in defining the energy bands. Various experimental techniques have been used to determine the density of localized states in the gap. It was concluded from transient photocurrentdecay measurements that the density of tail states decreases exponentially and does not have any defined structure [21]. It is now generally believed that, on top of a featureless distribution of states in the tails, a structured density of states exists, attributed to valence alternation pairs (VAPs)) [22]. For arsenic

1.3 Electronic Properties of Chalcogenide Glasses

7

triselenide, time-of-flight and transient photoconductivity measurements suggest a feature located 0.6 eV above the valence band edge which dominates the transport. There has been some controversy in the literature related to the sign of the correlation energy in amorphous selenium. While Kastner et al. [23] originally assumed a negative correlation energy, early theoretical work [24] indicated that, in fact, the correlation energy in a-Se should be positive, i.e. spin pairing at coordination defects would be energetically unfavorable. Recently, it was demonstrated by Kolobov et al. [25], using light-induced ESR, that the correlation energy was indeed negative. Popescu [26] has stated that in amorphous selenium the defect states C+ 3 (where C stands for chalcogen and the superscript and subscript gives the charge and coordination, respectively) give rise to discrete traps situated at around 0.33 eV below the bottom of the conduction band and they control the electron mobility. The C− 1 defect states are situated at 0.17 eV above the top of the valence band and control the hole concentration. Both types of traps are distorted by trapping of the charge carriers, and, as a consequence, their energies do not correspond to that found from light-absorption experiments and charge-carrier generation. Tanaka has recently proposed [27,28] a model of a realistic density of states based on several experimental results (Fig. 1.3a). Spatial potential fluctuations and the atomic structure in As2 S3 are shown schematically in Fig. 1.3b, c, respectively. In this model, it is assumed that the Urbach edge arises from fluctuations of the van der Waal’s type interlayer bonds and/or disordered interactions among the intralayer lone-pair electrons. The weak absorption tail in this model is ascribed to transitions involving antibonding states of As–As wrong bonds, which can produce unoccupied deep states below the conduction band. Electronic excitations can generate localized holes in the valence-band edge and delocalized electrons in the conduction band. The other possibility is generation of delocalized holes in the valence band and localized electrons in the conduction-band tail. Excited carriers are immediately (≈10−12 s) trapped into the localized states. S

S

As

Energy

As

As S

As

As

atomic distance

S

S

DOS

(a)

(b)

(c)

Fig. 1.3. (a) Density of states, (b) the spatial potential fluctuations and (c) the corresponding atomic structure proposed for As2 S3 glass (after [20])

8

1 An Introduction to Chalcogenide Glasses

1.3.2 Measurements of the Absorption Coefficient and the Optical Gap The optical absorption edge in amorphous semiconductors is generally not as steep as that in crystalline semiconductors. In general, the absorption spectrum α(¯ hω) can be divided into three parts [29] (see Fig. 1.4). For α ≥ 104 cm−1 , the spectrum shows a square-root dependence, α¯hω ∝ (¯ hω−EgT )1/2 .

4

10

2

absorption coefficient (cm−1)

10

0

10

−2

10

−4

10

1

2

3

photon energy (eV)

Fig. 1.4. Optical absorption edges in As2 S3 glasses at different temperatures. The two lines (solid and dashed ) show absorption spectra for different samples at 300 K. Spectra at 175 K(plus), 200 K (times), 250 K (downtriangle), 300 K (triangle), and 400 K (circle) have been obtained under 105 V cm−1 using the constant-photocurrent method. Also shown are spectra at 10 K for As2 S3 glass (dot-dashed line) and for a crystalline sample (dotted line) (after [30])

1.3 Electronic Properties of Chalcogenide Glasses

9

For 104 ≥ α ≥ 100 cm−1 , the so-called Urbach edge with the form of α ∝ exp(¯ hω/EU ) appears. For α ≤ 100 cm−1 , a weak-absorption tail with α ∝ exp(¯ hω/EW ) exists. In the above, EgT represents the optical (T auc) gap. In As2 S3 glass at room temperature, for example, EgT = 2.36 eV, EU ≈ 50 meV, and EW ≈ 250 meV [29–31]. It should be noted that the mobility gap, which can be evaluated from photoconduction spectra, appears to be located at ≈2.5 eV [30, 32]. The absorption coefficient, α, of chalcogenide films has been measured using several techniques for as-deposited, annealed, and photodarkened films, as well as in fabricated waveguides. At short wavelengths, where the absorption is high, a conventional spectrophotometer could be used, with corrections for reflection losses using the method described in [33]. At wavelengths beyond the band edge, however, the absorption is too small for this technique to be useful. Photothermal deflection spectroscopy (PDS) has therefore been used [31] to measure the relative absorption coefficient in the long-wavelength region. To calibrate the PDS data, absorption values from the spectrophotometer and the PDS are overlapped in the region of moderate absorption just beyond the band edge, where both give an accurate measurement. In addition, optical loss can be determined at spot wavelengths from propagation measurements made in slab waveguides. For these measurements [34], films deposited onto oxidized Si wafers were placed in a prism coupler, and the lowest-order slab waveguide mode was excited. Some radiation was scattered from the surface of the sample and could be detected using a cooled CCD camera. At wavelengths close to the absorption edge (633 nm), the decay of the intensity with distance was assumed to be dominated by film absorption, and hence the absorption coefficient for the film could be determined and used to calibrate the PDS data. The results of PDS measurements showed that as-deposited As2 S3 films have losses below 0.1 dB cm−1 across the telecommunications band at 1,300 and 1,550 nm [34]. Films that were 2.5 µm thick were used for single-mode waveguide fabrication using a direct-writing system. To assess the losses in fabricated waveguides, it was found possible to image the light scattered from the waveguides and to monitor the decay of the power in the waveguide as a function of distance using an IR-sensitive video camera. Light from laser-diode sources at 780, 1,300, and 1,550 nm was end-coupled into the waveguides using a microscope lens for these measurements. The losses obtained in this way were ∼0.4 dB cm−1 at 780 nm, ∼0.24 dB cm−1 at 1,300 nm, and 0.2 dB cm−1 at 1,550 nm [34]. These values are in good agreement with those obtained from the PDS measurements. The output from the waveguides was imaged with a microscope objective onto a video camera. The waveguides were single mode at 1,300 nm and 1,550 nm. Band theory for crystalline semiconductors suggests that the absorption coefficient for indirect transitions can be written as α = const × M 2

(hν − Eg )2 hν

(1.1)

10

1 An Introduction to Chalcogenide Glasses Table 1.1. Optical gap for evaporated films (after [35]) Eg (eV )

composition

As2 S3 2.26 1.935 As3 Se97 1.872 As22 Se78 1.79 As41 Se59 1.79 As39 Se61 1.922 Ge9 As25 Se66 1.828 Ge5 As35 Se60 1.783 Ge2 As40 Se58 The optical gap for As2 S3 is for PLD deposited films [34]

Table 1.2. Optical gap for some bulk chalcogenide glasses (after [36]) glass

Eg (eV )

Ge25 Se75 Ge25 Se65 Te10 Ge28 Se60 Sb12 As2 Se3 As2 S3 The optical gap for As2 S3 is

2.07 1.73 1.8 1.77 2.4 taken from [37]

where M is the matrix element of the optical transition and Eg is the band gap energy. The absorption in many amorphous semiconductors is observed to obey this relation above the exponential Urbach edge. If M is constant, plotting (αhν)1/2 versus hν should result in a straight line. The optical gap, Eg , is obtained from the intersection of this line with the energy axis. The value obtained for the optical gap of pulse laser deposited (PLD) a-As2 S3 film was Eg = 2.26 ± 0.02 eV [34]. This value is slightly lower than that of 2.36 eV found for thermally evaporated As2 S3 films, most probably due to illumination of the films during the laser-deposition process. Table 1.1 shows the results of optical-gap measurements for evaporated chalcogenide films, while the optical gaps of some chalcogenide glasses are shown in Table 1.2.

1.4 Chalcogenide Glasses for Near-Infrared Optics A range of optical functions can be realized in these glasses, including optical amplification and emission at telecommunication wavelengths by rare-earth (e.g., Pr and Er) dopants, fabrication of waveguides (channel, self-written, femtosecond written), that can laterally or vertically couple light to various locations within a planar structure, and gratings (relief and phase) that can spectrally filter or modulate light. Diffraction gratings have been fabricated

1.4 Chalcogenide Glasses for Near-Infrared Optics

11

in chalcogenide glasses using the photoinduced effects that they exhibit. Both photodarkening [38] and the metal-photodissolution effect [40] (especially of silver) have been used to fabricate transmissive gratings, especially for use at IR wavelengths. A variety of techniques have been used to fabricate these gratings, including holographic, mask exposure, or etching methods. These gratings can be used as efficient beam combiners, couplers and have significant applications in monochromators, laser-tuning devices, shapers, opticalfiber couplers, etc. For instance, gratings have been written holographically in sputtered Ge10 As40 S25 Se25 films with 514.5 nm light from an Ar-ion laser and probed with a 670 nm diode laser [38]. The probe was highly attenuated to avoid influencing the photodarkening process. The first-order efficiency of Raman–Nath diffraction for a probe beam tuned to the Bragg condition was used to measure a refractive-index change of ∆n = 0.001 for a written intensity of 100 mW cm2 . Slinger et al. [39] recorded volume holographic gratings in evaporated As40 S60 films in which silver was photodissolved. They measured the angular response of the gratings, and replay was made in air using light of wavelength 632.8 nm from a helium–neon laser. Typical Bragg behavior was observed and the diffraction efficiency in the first-order diffracted beams reached a maximum of 9% near the Bragg angle. A photolithographic technique has been used [40] to fabricate surface-relief gratings in a bi-layer structure consisting of a 0.8 µm a-As30 S70 film and an under-layer of Ag of 0.14 µm thickness. Thin gratings produced by mask exposure showed Raman– Nath type diffractive behavior and, when replayed at 632.8 nm, first-order diffraction efficiencies of up to 10% were measured. First- and second-order Bragg reflectors at telecommunication wavelengths (1.5 µm) have been fabricated in single-mode monolayer (As2 S3 ) and multilayer (As–S–Se/As–S) chalcogenide glass planar waveguides with near-band gap illumination using an interferometric technique [41]. Reflectivities as high as 90% near 1.55 µm, and refractive-index modulations up to 3 × 10−4 were achieved. The volume photodarkening effect is the principal mechanism involved in the formation of the Bragg gratings. The stability and high efficiency of these gratings make them potentially useful as wavelength-selection elements, and add-drop filters for WDM networks [41]. Richardson et al. [42] have written permanent waveguides in both bulk and film of As2 S3 glasses [43]. Using a train of 850 nm femtosecond laser pulses, they measured both the induced index variation and structural changes induced through the photomodification. The refractive-index variation between the waveguide (exposed region of the glass sample) and the cladding (unexposed region) was evaluated following waveguide writing. An induced index change of ∆n = −0.04 was associated with the formation of a 9 mm diameter circular waveguide formed by moving the sample through a well-characterized focal region [41]. A key finding of the study defined the structural mechanism associated with the writing process in the As2 S3 material. The concurrent destruction of As–S bonds within the glass network and the associated formation of As–As bonds during the bulk material

12

1 An Introduction to Chalcogenide Glasses

modification was quantified by a two-dimensional micro-Raman analysis (excitation with λ ≈ 752 nm). Optical amplification at 1.083 µm in neodymium-doped chalcogenide fibers was observed [44] in a glass composition of Ge–As–Ga–Sb–S. A maximum internal gain of 6.8 dB was achieved for a pump power of 180 mW. The first amplified spontaneous emission in a chalcogenide glass fiber has also been reported [44]. Laser action in a rare-earth-doped GaLaS chalcogenide glass has been demonstrated, showing that this class of glasses is suitable for active applications, such as amplifiers and lasers [45]. A neodymium-doped GaLaS glass laser has been operated under continuous-wave conditions at a wavelength of 1.08 µm when pumped with a Ti:Sapphire laser at either 0.815 or 0.890 µm [45]. The reasonably low laser threshold indicated acceptable glass losses, but the laser performance was worse in comparison with conventional Nd-lasers. The original application of GaLaS glass was as a practical and efficient 1.3 µm optical fiber amplifier. The ability to move and manipulate light without using costly and delicate articulated arms and mirrors is a common requirement for several applications, e.g., laser surgery. Power handling of GaLaS fibers has been assessed by coupling to an Nd:YAG laser operating at 1,064 nm. A total of 5 W of power was guided through a core of about 150 µm with no apparent laser damage [46].

1.5 Chalcogenide Glasses for Mid-IR and Far-IR Applications A selenide glass has been developed that can be doped with rare-earth ions and is stable against crystallization during fiberization [47]. The glass is based on GeAsGaSe and can be doped with Pr3+ and Dy3+ for near- and mid-IR applications. The doped glasses have been fiberized with core-only losses of 0.8 dB m−1 at 6 µm and 1.5 dB m−1 at 2.5 µm. Single-mode fibers have been drawn with a measured core loss of 3 dB m−1 at 1.55 µm. Pr3+ incorporation has been investigated and mid-IR emission in the 3–5 µm region has been observed. Schweizer et al. [48] have found that the absorption spectrum of a 9.7 mol% Er3+ -doped GaLaS glass showed excellent rare-earth solubility and the potential for high doping concentrations and hence short devices; this represents a major advantage of GaLaS glass compared with conventional chalcogenide glasses, which suffer from very low rare-earth solubilities. Values obtained for branching ratios of 1% for both the 3.6 and 4.5 µm transitions, with measured lifetimes of 100 and 590 µs and cross-sections of 0.43 × 10−20 and 0.25 × 10−20 cm2 , were in good agreement with the results of [49], respectively, in an Er3+ -doped barium indium gallium germanium sulfide glass. The radiative properties of Er3+ -doped Ga– La–S lend themselves to applications. Radiation at 2 µm has application in LIDAR systems, radiation at 2.75 µm coincides with a strong water absorption

1.5 Chalcogenide Glasses for Mid-IR and Far-IR Applications

13

in tissue and is used for medical applications, the 3.6 µm transition could be useful for H2 S, NO, and SO2 (remote) sensing and the 4.5 µm transition could find use in CO and O3 gas sensors when tuned to 4.7 µm. Highpower CO and CO2 lasers operating at 5.4 and 10.6 µm, respectively, are available and are used for industrial welding and cutting. Transmitting the laser power through fibers enables remote operation to take place. Te-based fibers have demonstrated output powers of 10.7 W for 19.4 W launched power (efficiency = 55.2%) at 10.6 µm [50]. The fibers possessed an antireflection (AR) coating and were cooled with water to prevent thermal lensing caused by an increase in absorption coefficient with temperature (dα/dT ) and an increase in refractive index with temperature (dn/dT). On the other hand, arsenic sulfide-based fibers have demonstrated 85 W output power for 169 W launched power (efficiency = 50.3%) without the need for cooling and AR coatings [51]. Unlike the Te-based glasses, arsenic sulfide-based glasses have smaller values of dα/dT and dn/dT . Typically, the fiber diameters are usually in excess of 500 µm for high-power laser delivery to reduce the power density. However, small core diameter ( 0 (after the focus), however, the positive lensing causes the beam divergence to decrease, resulting in an increased aperture transmittance. The lens has little effect on the beam near z = 0, and the aperture transmittance returns to its low-intensity value. The net Z-scan yields an s-shaped transmittance curve (see Figs. 3.7 and 3.8). A material with a negative n2 will produce a similar curve, but with the peak and valley reversed about z = 0. The basis of the Z-scan technique is the fact that the aperture transmittance as a function of sample position depends on the magnitude and sign of n2 . The nonlinear medium impresses a phase distortion on the electric field of the transmitted light. The value of n2 can be extracted using the computed aperture transmittance. Among several advantages of the Z-scan technique is its simplicity, and being a single- beam technique, there are no difficult alignment problems other than keeping the beam centered on the aperture. It can be used to determine both the magnitude and sign of n2 . The sign is obvious from the shape of the transmittance curve. Unlike most DFWM methods, the Z-scan can determine both the real and the imaginary parts of χ(3) . The Z-scan technique is also highly sensitive, capable of resolving a phase distortion of ≈λ/300 [108] in samples of high optical quality. Among the disadvantages of the technique is the requirement of a high-quality Gaussian

Fig. 3.7. Transmission versus position, z, for a closed- aperture scan of Ge33 As12 Se55 at an intensity of 0.16 GW cm−2 (after [118])

3.5 Third-Harmonic Generation

63

Fig. 3.8. Transmission versus position, z, for an open-aperture scan of Ge33 As12 Se55 at an intensity of 0.16 GW cm−2 (after [118])

TEM00 beam for absolute measurements. Effects such as sample distortion and tilting of the sample during translation can cause beam deflection and hence cause unwanted fluctuations in the detected signal.

3.5 Third-Harmonic Generation Generally, one utilizes a Q-switched pulse Nd:YAG laser for this technique, which provides nanosecond pulses at low repetition rates (10–30 Hz). After proper selection of wavelength and polarization, the laser beam is divided into two parts, one being used to generate the third harmonic in the sample and the other to generate the third harmonic in a reference (see Fig. 3.9). Fused silica glass is generally taken as the reference. The third-harmonic generation (THG) susceptibility is determined from the following relation. χ(3) (THG) =

n+1 ns + 1

4

lc.s lc



I3ω I3ω,s

1/2 χ(3) s (THG).

(3.14)

Here n is the refractive index, lc is the coherence length, and I3ω is the TH intensity. The suffix s means the standard medium. For nonphase-matched THG, one uses the Maker fringe method in which the path length of the sample is varied and the third-harmonic signal is monitored as a function of the interaction length to obtain the fringes. From the fringes, one determines the coherence length for the sample. Since all media, including air, show thirdorder nonlinear effects, one has to be extremely careful in third-harmonic

64

3 Experimental Techniques to Measure Nonlinear Optical Constants 0.68 mm Difference Frequency generator

Dye laser

Q-switch YAG laser

SHG

trigger

1.06 mm Driver

1.9 mm

Rotator Lens

w

Boxcar averager

Computer

Filter

H.V.

3w P.M.T

Pump Sample

THG

Fig. 3.9. Schematic illustration of the equipment for measurement of third-harmonic generation. SHG, second harmonic generation; THG, third harmonic generation; P.M.T, photomultiplier tube; H.V., high voltage (redrawn from [119])

measurements. Contributions from air and the walls of the cell (the latter in the case of liquid samples) may even be dominant, especially with samples having low values of χ(3) . The resulting harmonic field is equal to the sum of contributions from consecutive nonlinear media. The advantage of the THG technique is that it probes purely electronic nonlinearity. Therefore, orientational and thermal effects, as well as other dynamic nonlinearities derived from resonant excitations, are eliminated. The disadvantage of the technique is that very large resonant dynamic nonlinearities cannot be probed by this method. The THG method does not provide any information on the time response of optical nonlinearity.

3.6 Optical Kerr Gate and Ellipse Rotation 3.6.1 Optical Kerr Gate In this method, an intense linearly polarized light pulse traveling through an optically isotropic χ(3) medium induces optical birefringence using the optical Kerr effect. A linearly polarized weaker probe pulse is utilized to obtain the birefringence δn = δn −δn⊥ . One can obtain χ(3) from δn. The time evolution of the birefringence, and hence the response time of χ(3) , can be probed by delaying the probe beam with respect to the pump beam. A suitable Kerrgate experiment is shown in Fig. 3.10 [120]. A picosecond pulse of appropriate wavelength is divided in two parts: a strong (100 MW cm−2 ) pump pulse and

3.6 Optical Kerr Gate and Ellipse Rotation

65

C.C Pump P

A

S

Probe M1

P M2 Laser

M3 B

Fig. 3.10. Schematic of the Kerr-gate experimental arrangement. A, analyzer; P, polarizer; S, sample; M, mirrors; B, beam splitter; C.C., corner cube for optical delay (redrawn from [120])

a weak probe pulse that undergoes a variable delay. The two beams make an angle (usually 45◦ ) with each other. The analyzer measures the transmitted probe beam as a function of the delay time between the probe and pump pulses. The signal It (τ ) transmitted by the analyzer, for a time delay τ , is given by:    +∞ δφ(t) 2 Eprobe (t − tD ) sin2 It (τ ) = dt. (3.15) 2 −∞ Here Eprobe is the electric field of the probe beam and δφ is the phase retardation of the probe beam due to induced birefringence given by 2πl δn(t), (3.16) λ where l is the sample path length, λ is the wavelength of the probe beam, and δn(t) is the induced refractive-index change written as

 ns  t t − t f 2i 2 2  δn(t) = n2 E1 (t) + E1 (t ) exp − (3.17) dt . τ τ i i −∞ i δφ(t) =

Here, nf2 is the intensity-dependent refractive index due to the rapidly responding electronic χ(3) . The second term consists of various slowly responding nonlinearities with response time τi . In the case of nonresonant electronic nonlinearity δn = n2 Ipump .

(3.18)

If δn is substituted in (3.16), the peak value of the phase retardation δφ is obtained from 2πl δφ = n2 Ipump . (3.19) λ

3 Experimental Techniques to Measure Nonlinear Optical Constants 0.8

CS2 reference (1 mm)

(a)

0.6 0.4 0.2 0.0 −800 −400

0

400 800 1200 1600

Optical Kerr signal (a.u)

Optical Kerr signal (a.u)

66

16 14 12 10 8 6 4 2 0

glass sample (0.7 mm)

(b)

150 fs

−600 −400 −200 0

200 400 600

Delay time (fs) Fig. 3.11. Time-resolved optical Kerr signals at 820 nm. (a) Standard CS2 reference medium. (b) The 90GeS2 − 5Ga2 S3 − 5CdS (in mol.%) chalcogenide glass sample. Dots show the experimental data and solid curve indicates the Gaussian fit (after [121]) Reprinted from X.F. Wang, Z.W. Wang, J.G. Yu, C.L. Liu, X.J. Zhao, Q.H. Gong, c (2004), with permission from Elsevier Chem. Phys. Lett. 399 (2004) 230, 

From a measurement of the peak value of the transmitted probe signal Is , one can obtain δφ and hence n2 , which is related to χ(3) by n2 =

12π (3) (3) (χ − χ1122 ). n0 1111

(3.20)

At a wavelength far from resonance and for an isotropic medium, a purely (3) (3) electronic nonlinearity leads to χ1122 = 13 χ1111 and hence the measurement of (3) n2 yields a direct determination of χ1111 . A typical optical Kerr-gate result for a 90GeS2 − 5Ga2 S3 − 5CdS sample is shown in Fig. 3.11.

3.6.2 Ellipse Rotation If a single optical beam is incident on an isotropic nonlinear medium, for both linearly and circularly polarized light, the third-order polarization has the same vector character as the applied field. Thus, the induced birefringence produces no change in the polarization state of the optical field. When an elliptically polarized wave induces a nonlinear polarization that mixes the left- and right-hand circularly polarized components of the wave, an induced circular birefringence ∆nc is formed and is given by ∆nc = −

(3)  3χxyyx  2 2 |E+ | − |E− | , n0

(3.21)

where E+ and E− are the complex field amplitudes of the left- and righthand circularly polarized components, respectively, of the elliptically polarized

3.7 Self-Phase Modulation

67

wave. The circular birefringence produces a rotation of the polarization ellipse (ellipse axes) through an angle θ, given by θc =

π∆nc L . λ

(3.22) (3)

A measurement of this rotation angle gives a measure of the χxyyx tensor component of the medium.

3.7 Self-Phase Modulation An intense pulse propagating through a nonlinear medium acquires an additional phase due to the nonlinear index of refraction, n2 . The self-induced phase is time dependent if the pulse intensity is time dependent. The nonlinear phase shift can be written as φNL (t) = −

ω0 I n I(t)L, c 2

(3.23)

where ω0 is the pulse center frequency and L is the medium thickness. An instantaneous frequency shift is introduced: δω(t) = dφNL /dt.

(3.24)

The frequency shift near the peak of the pulse is zero. The leading edge is red-shifted while the trailing edge is blue-shifted. If the bandwidth of the pulse is τ0−1 , self-phase modulation becomes important when ∆φmax ≥ 2π, where ∆φmax is the maximum phase shift. For light with a wavelength 1 µm propagating through a medium of length 1 cm with nI2 ≈ 10−18 m2 W−1 , a peak intensity of >10 GW cm−2 is needed, which means pulses of picosecond width and shorter are required. A self-modulated pulse has a broadened spectrum of frequencies. If the medium has group-velocity dispersion (for ultrashort pulses), then the pulse will also spread in time [122]. Applications of selfphase modulation include super-continuum generation [123] and pulse compression [122]. A mode-locked Nd:YAG laser is normally used as the light source [124] (see Fig. 3.12). The temporal-pulse profile is monitored with a pin-photodiode and a sampling oscilloscope. The laser beam is focused into the fiber with an objective lens. Spectral broadening of the output is analyzed using an optical spectrometer. The n2 value can be estimated from the following equation:

4 ∆ν 2 = ∆ν02 1 + √ γ 2 P 2 , (3.25) 3 3 where γ=

2πn2 Zeff λAeff

(3.26)

68

3 Experimental Techniques to Measure Nonlinear Optical Constants Sampling oscilloscope 1.319 mm mode-locked Nd:YAG

PD

Camera

ATT C

A Power meter or spectrometer l/2

L

fiber

L

L

L

M

P

Fig. 3.12. Schematic diagram of the experimental set up. PD, pin-photodiode; ATT, attenuator; C, chopper; λ/2, half-wave plate; L, lens; A, variable aperture; M, glass plate; P, polarizer (redrawn from [124])

NORM. SPECTRAL INTENSITY

Reprinted from M. Asobe, K. Suzuki, T. Kanamori, and K. Kubodera, Appl. Phys. c (1992) with permission from the American Institute of Lett. 60, 1153 (1992),  Physics 1 High Intensity Low Intensity

0.8

0.6

0.4

0.2

0 1550

1555

1560

1565

1570

1575

WAVELENGTH (nm) Fig. 3.13. Self-phase modulation spectrum at the output of an As2 S3 waveguide. The solid and dashed lines are for high and low input intensity, respectively (after [125]) Reprinted from K.A. Richardson, J.M. McKinley, B. Lawrence, S. Joshi, A. Vilc (1998) with permission from Elsevier leneuve, J. Opt. Mater. 10, 155 (1998), 

and

1 [1 − exp(−αL)]. (3.27) α Here, ∆ν0 is the initial spectral width, P is the input peak power, λ is the wavelength, Aeff is the effective core area, Zeff is the effective length, α is Zeff =

3.8 Spectrally Resolved Two-Beam Coupling

69

the loss coefficient, and L is the fiber length. A typical result of self-phase modulation for an As2 S3 waveguide is shown in Fig. 3.13.

3.8 Spectrally Resolved Two-Beam Coupling Spectrally resolved two-beam coupling (SRTBC) is a two-beam technique and it utilizes the modulation of the probe-pulse spectrum due to the nonlinear phase shift induced by the pump [126]. SRTBC provides the sign and magnitude of the real and imaginary parts of the third-order susceptibility (nonlinear refraction and nonlinear absorption) along with their dynamics. This technique is based on a standard pump-probe setup with the addition of a monochromator to analyze spectrally the energy content of the probe beam (see Fig. 3.14). The presence of a strong pump pulse can change the index of refraction and absorption coefficient of a material. When the weak probe beam passes through the medium, it acquires a frequency shift and may experience nonlinear absorption. These changes in the probe beam are detected by measuring the energy of the beam in a narrow spectral band selected by the monochromator as a function of the time delay between the arrival of the pump and probe pulses. One can extract the nonlinear index of refraction and two-photon absorption coefficient from these data. Since the experiment monitors changes in the spectrum, the signal is derivative-like; an index change with a Gaussian temporal envelope will produce a bimodal signal. The TPA signal is superimposed on this and is directly proportional to the energy loss. The advantage of the technique is detection far away from the center wavelength. At this wavelength, the intensity of the beam is much smaller than at the center of the spectrum, but the nonlinear modulation is relatively larger. Nonlinear phase shifts as small as 10−6 rad can be detected. The time resolution and sensitivity of the technique is such that in the case of fused silica, which has among the smallest known nonlinearity, the nuclear contribution, which is only a fraction

Probe

Pump

Sample

l

Detector

Delay

Fig. 3.14. Schematic diagram of the experimental set-up for a spectrally resolved two-beam coupling (SRTBC) technique. λ represents the monochromator

70

3 Experimental Techniques to Measure Nonlinear Optical Constants

∆T / T (10−3)

2 1

Spectrum (A.U)

3 measurement calculation

1.48 1.52 1.56 1.60 Wavelength (mm)

0 −1 −2 −400

−200

0

200

400

Delay (fs) Fig. 3.15. Representative SRTBC signal obtained with an Er-doped fiber laser at 1.25 µm (EDFL). The calculated signal is derived from a zero-phase Fourier transform of the laser spectrum (inset). The difference between measurement and calculation for positive delay is due to the nuclear contribution to the nonlinearity (after [128])

of the total nonlinearity, is clearly resolved [127]. A typical result of SRTBC is shown in Fig. 3.15.

3.9 Mach-Zehnder Interferometry The experimental set up is shown in Fig. 3.16. As is seen, the first arm of a Mach-Zehnder interferometer is used as the reference beam I1 (r). The second arm I2 (r) is used as the probe beam. The nonlinear medium (NS) is illuminated by a pump beam focused by means of the lens L1 . At the output of the set-up, the interferometer pattern intensity INL (r) is recorded on a CCD camera placed in a O(x, y) plane perpendicular to the light-propagation axis. Rectilinear fringes are obtained by adjusting the interferometer. The intensity at the output is given by  (3.28) INL (r) = I1 (r) + I2 (r) · T (r)2 + 2T (r) I1 I2 cos[φL + φNL (r)], where φL is the linear phase difference between the interferometer arms. Local fringe alterations occur in a small region of the interference pattern. These local alterations are due to a local displacement of fringes attributed to the nonlinear dephasing φNL (r) and a local modification in the fringe visibility attributed to the amplitude nonlinear transmission T (r). A numerical spatial Fourier transform (FT) is performed on the acquired image INL (r). The FT

3.9 Mach-Zehnder Interferometry

71

Ic ( r ) = control

M

CCD

BS L I1 ( r ) = reference L I2 ( r ) = Probe

M BS

x M NS

M

z y

L1 Pump NS: nonlinear sample; L: lens; M: mirror; BS: beam splitter

BS

Fig. 3.16. Set-up for Mach-Zehnder interferometry measurements (redrawn from [129, 130])

of the cosine function (representing the rectilinear fringes) in the third term of (3.28) consists of two Dirac delta functions far enough from the origin (location of the FT of the first and second terms in (3.3)–(3.28) in the Fourier plane), convoluting a complex function that contains the information about the nonlinearities. By considering the part of the spectrum around one of these delta functions and performing an inverse FT on this part, the quantities T and φNL as functions of the radial coordinate r are extracted. In order to extract the information related to the nonlinearities, a linear fringe pattern is obtained by placing the sample in the interferometer but the pump beam is switched off. In this case, the nonlinearities are negligible, which allows one to measure the linear response of the sample. By comparing the linear and the nonlinear fringe patterns, one is able to get spatially resolved information on the nonlinearities. The nonlinear dephasing allows one to find a nonlinear index assuming a purely third-order nonlinearity. φNL (r) = kLn2 Ieff (r), with Ieff (r) =

1 ln[1 + q(r)], βL

(3.29)

(3.30)

72

3 Experimental Techniques to Measure Nonlinear Optical Constants

where Ieff is the effective pump intensity inside the nonlinear medium, k is the wave vector, L is the sample length, β is the two-photon absorption coefficient −αL and q(r) = βLeff (r) with Leff = (1−eα ) . I(r) is the incident pump-intensity distribution and α is the linear absorption coefficient. The two-photon absorption coefficient β is deduced from the amplitude transmittance as T (r) = (eαL [1 + q(r)])1/2 .

(3.31)

Figure 3.17 shows a typical result of Mach-Zehnder interferometry for an As2 S3 sample. (b) y

250

50

200

50

200

100

150

150

100

200

50

100

150

150

100

200

50

Intensity (a.u.)

250

250 x

50

250

100 150 200 250

(d) y 0.6 50

0.4

100

0.2

150

0

200

−0.2

250

−0.4 x

50

100 150 200 250

Nonlinear dephasing (rad.)

(c)

50

x

10−4

1 ∆n 0.8

Intensity (a.u.)

(a) y

100 150 200 250 ∆n versus incident intensity

4096 dots

0.6 0.4 0.2 00

As2S3 n2 = 5 10−18 m2 / W 5

10 15 I (W / m2)1012

Fig. 3.17. Experimental acquisition and numerical processing in the presence of a chalcogenide glass As2 S3 (1.8 mm thick) at λ = 1064 nm. X and y are graduated in pixels. The control and the pump beams are not shown in the figure. (a) The fringe pattern without the incident pump beam. The fringes that appear in this figure are due to interference between reference and probe beams. (b) Local alteration of the previous fringe pattern in the presence of the incident pump beam inducing nonlinear dephasing. (c) Result of the numerical processing giving an image of the nonlinear dephasing when the two experimental acquisitions shown in (a) and (b) are taken into account. (d) Plot of ∆n(x, y) versus I(x, y). The n2 measurement is given by the slope of the linear regression line calculated over 4096 pixels in the image (after [129]) Reprinted from G. Boudebs, F. Sanches, J. Troles, F. Smektala, Opt. Commun. 199 c (2001), with permission from Elsevier (2001) 425, 

3.10 Summary

73

3.10 Summary Different techniques have been introduced, which allow measurements of the nonlinear optical constants (n2 and β) of bulk/film samples. While DFWM allows determination of n2 only, Z-scan can be used to obtain both n2 and β of samples. Moreover, this single-beam technique allows determination of the sign of n2 . THG is a technique that probes purely electronic nonlinearity but does not provide any information on the time response of optical nonlinearity. An intense pulse propagating through a nonlinear medium acquires an additional phase due to the nonlinear index of refraction. By monitoring the temporal-pulse profile, and analyzing the spectral broadening of the output due to self-phase modulation, the n2 value of the sample can be estimated. SRTBC is a two-beam technique that provides the sign and magnitude of the real and imaginary parts of the third-order susceptibility along with their dynamics. Finally, Mach-Zehnder interferometry allows one to determine both n2 and β. When the nonlinear medium is illuminated by a focused pump beam, local fringe alterations occur in a small region of the interference pattern. These local alterations are due to the nonlinear dephasing φNL (r) and the amplitude nonlinear transmittance T (r). The quantities T and φNL as functions of the radial coordinate r are extracted. The nonlinear dephasing allows determination of the nonlinear index, while the two-photon absorption coefficient β is deduced from the amplitude transmittance.

This page intentionally blank

4 Measurement of Nonlinear Optical Constants

4.1 Measurements of Nonlinear Refractive Index n2 Different techniques, such as two-photon absorption spectroscopy [131], degenerate four-wave mixing (DFWM), Z-scan [132, 133], third-harmonic generation (THG), optical Kerr-shutter (OKS) [134], and self-phase modulation (SPM), have been used to measure the nonlinear refractive index, as well as the nonlinear absorption coefficient, of chalcogenide glasses. The nonlinear refractive index n2 of As2 S3 glass fiber was first measured directly by a spectrum-broadening experiment at a wavelength of 1.3 µm [135]. The n2 value was also measured by a Kerr-shutter (Fig. 4.1) experiment at 1.3 µm [136]. In the optical Kerr effect, the nonlinear phase shift induced by an intense, high-power, pump beam is used to change the transmission of a weak signal through a nonlinear medium. The operating principle of a Kerrshutter can be explained in the following way. The pump and signal beams are linearly polarized at the fiber input with a 45◦ angle between their directions of polarization. A gate pulse induces a phase-shift difference between two signal components whose polarization is parallel and perpendicular to that of the gate. When the phase-shift difference reaches π, the polarization of the signal is switched by 90◦ . The n2 values obtained were 1.7 × 10−14 cm2 W−1 and 4×10−14 cm2 W−1 , respectively. The difference may be due to an increase in the transmission loss caused by a higher average power in the spectrumbroadening experiment. Other Kerr-shutter experiments at 1.55 µm resulted in n2 = 2 × 10−14 cm2 W−1 . There is a possibility that the n2 value depends on the wavelength because of the two-photon resonance near 1.3 µm [138]. So, it was confirmed that As2 S3 glass possesses a n2 value about two orders of magnitude greater than silica (n2 = 3 × 10−16 cm2 W−1 ). The χ(3) value obtained for the Kerr-shutter experiment at 1.3 µm was 1.4 × 10−19 m2 V−2 and was in good agreement with the value of 1.01 × 10−19 m2 V−2 measured in a THG experiment at 2 µm [139]. Third-order optical nonlinear susceptibilities χ(3) of some high-refractive-index chalcogenide glasses were evaluated from

76

4 Measurement of Nonlinear Optical Constants

Fig. 4.1. Principle of an OKS (after [137]) c (1997), with Reprinted from M. Asobe, Opt. Fiber Technol. 3 (1997) 142,  permission from Elsevier

Fig. 4.2. Relationship between linear and third-order optical susceptibilities (after [137]) c (1997), with Reprinted from M. Asobe, Opt. Fiber Technol. 3 (1997) 142,  permission from Elsevier

THG [139]. Compared with oxide glasses, whose χ(3) was known, χ(3) values of chalcogenide glasses were higher by an order of magnitude. The highest χ(3) (for a composition of As40 S57 Se3 ) is 1.96 × 10−19 m2 V−2 , being comparable with those of high-χ(3) organic compounds (see Fig. 4.2). It was found [133] that χ(3) generally increased with increasing density of the chalcogenide glasses. The real part of χ(3) of binary and ternary glasses is plotted against Eg and n in Fig. 4.3. Re χ(3) increases monotonically with decreasing Eg and increasing n. A similar relation holds between Im χ(3) and Eg and n. As a result, the χ(3) of binary and ternary glasses derived from the Z-scan method increases monotonically with decreasing Eg and increasing n. This agrees well with the dependence of χ(3) on Eg and n measured by THG [140]. It should be mentioned here that we have tried to give χ(3) values in MKS units

4.1 Measurements of Nonlinear Refractive Index n2 3

3 (a)

Re χ(3) / 10−11 esu

77

(b)

2

2

1

1

0 2.0

2.5

3.0

Eg / eV

3.5

0 2.0

2.2

2.4

2.6

n

Fig. 4.3. The real part of χ(3) of binary La2 S3 –Ga2 S3 and ternary MS–La2 S3 –Ga2 S3 glasses plotted against (a) Eg and (b) n, (MS = Ag2 S and Na2 S) (after [133])

throughout the text whenever possible. In the literature, however, χ(3) values are normally given in Gaussian or (esu) units. The nonlinear susceptibilities in these two systems of units are related by χ(3) (MKS) =

4π χ(3) (Gaussian) = 1.40 × 10−8 χ(3) (Gaussian), (3 × 104 )2

where the unit of χ(3) in the MKS system of units is m2 V−2 . Harbold et al. [141] have found that chalcogenide glasses in the As–S–Se system simultaneously exhibit a large nonlinear index of refraction and a figure of merit (FOM = n2 /βλ) that satisfies a standard criterion for alloptical switches (AOSs) [142]. They found that in samples with a FOM >5, nonlinear phase shifts of π rad can be produced without damage for intensities of ≤200 MW cm−2 . They observed that the FOM increases substantially near hν/Egap ≈ 0.45, as is qualitatively expected when the absorption edge is not infinitely sharp. Their experiments confirm that the nonlinearities are determined largely by the abundance of the most polarizable constituent, in this case selenium. In particular, As40 Se60 exhibits high values of n2 and FOM (2.3×10−17 m2 W−1 and 11, respectively) and is thus quite promising for AOS use at 1.55 µm. A monotonic increase in n2 with progressive replacement of S by Se in the sulfoselenide glasses, and the gradual removal of Se at fixed Ge:As ratio from the selenide glasses, is observed [143]. As shown in Fig. 4.4, n2 increases faster than β for photon energies just below the half gap. The FOM (Fig. 4.5) depends on the proximity of the frequency of the light to the two-photon absorption (TPA) edge and the peak at hν/Egap ≈ 0.45 occurs when nonlinear refraction increases more rapidly than TPA with normalized photon energy. Quemard et al. [144] suggest that the concentration

78

4 Measurement of Nonlinear Optical Constants

Fig. 4.4. Variation of n2 and β, the TPA coefficient, with normalized photon energy in the Ge–As–S–Se system (after [143]) Reprinted from J.M. Harbold, F.O. Ilday, F.W. Wise, and B.G. Aitken, IEEE c (2002) with permission from IEEE Photon. Technol. Lett. 14, 822 (2002), 

Fig. 4.5. Variation of n2 with normalized photon energy in the selenide, sulfoselenide, sulfide, and heavy-metal-doped oxide systems, all at 1.25 µm (after [3]). Fused silica is also shown for comparison. Glasses in [4] in the inset belong to Ge–Se and Ge–Se–As systems Reprinted from J.M. Harbold, F.O. Ilday, F.W. Wise, and B.G. Aitken, IEEE c (2002) with permission from IEEE Photon. Technol. Lett. 14, 822 (2002), 

4.1 Measurements of Nonlinear Refractive Index n2

79

Fig. 4.6. Variation in the FOM with normalized photon energy in the Ge–As–Se and Ge–As–S–Se systems. Two femtosecond laser sources at 1.25 and 1.55 µm were used in nonlinear measurements using SRTBC (after [143]) Reprinted from J.M. Harbold, F.O. Ilday, F.W. Wise, and B.G. Aitken, IEEE c (2002) with permission from IEEE Photon. Technol. Lett. 14, 822 (2002), 

of electron lone pairs is the dominant factor in achieving large nonlinearities, and it has recently been confirmed that n2 increases with the most polarizable constituent (in this case, selenium) in As–S–Se glasses [141]. It should be noted that a change in the electron lone-pair concentration modifies the energy gap so it is impossible to determine whether a corresponding change in the nonlinearity is due to lone pairs or resonant enhancement. However, Harbold et al [143] found no systematic increase of n2 with increasing Se content and hence lone-pair concentration. Overall, they found that n2 increases as the band gap decreases in the selenide and sulfoselenide glass systems. It can be said that the general trend in the nonlinearity is accounted for by the normalized photon energy (Fig. 4.6). So, when designing an AOS at a given wavelength, one would choose a glass with hν/Egap ≈ 0.45 in order to achieve both a large n2 and large FOM. Kosa et al. [145] measured the third-order nonlinear optical response of silver-doped and undoped As2 S3 at a wavelength of 1.064 µm. A single-beam (Z-scan) technique was used for the measurement, which allowed them to separate the refractive and absorptive contribution to the signal at this wavelength. They observed that silver doping changes the sign of the refractive nonlinearity and it was also found that n2 increased by almost a factor of 80 relative to the undoped material. Third-order nonlinear optical properties of chalcogenide glasses were investigated by Kanbara et al. [132] through THG, OKS, and DFWM measurements. They examined the dependence of the THG

80

4 Measurement of Nonlinear Optical Constants

susceptibility on the absorption edge (for glasses in As–S–Se and Ge–As–S systems), showing that the susceptibility rapidly increased as the absorption edge red shifted. To within experimental errors, the THG susceptibility compared well with the OKS susceptibility. The DFWM experiment pointed out that an ultrafast response time of less than a picosecond was attainable with the chalcogenide glass. The Z-scan measurements were performed on germanium arsenic selenide films of 2 µm thickness [146]. 100-fs pulses with energies in the range of 0.1–0.5 µJ were used. In most measurements, the focused spot size was in the range of 20–40 µm, which resulted in maximum light intensities in the range of 10–150 GW cm−2 . A simple arrangement allowed the open-aperture Z-scan and the closed-aperture Z-scan to be recorded simultaneously. The Z-scans obtained were analyzed with expressions derived by Sheik-Bahae et al. [138] to yield the real part of the nonlinear phase shift ∆φreal induced by the thirdorder nonlinearity and the Tt factor (defined here as Tt = 4π∆φimag. /∆φreal ) for a given sample. This analysis was performed by comparing the shapes of closed- and open-aperture scans with those computed theoretically. Roughly speaking, the amplitude of a closed-aperture Z-scan (i.e., the peak-to-valley difference in transmission values) is proportional to the real part of the nonlinear phase shift ∆φreal , whereas the asymmetry of a closed-aperture scan depends on the Tt factor (for Tt = 0, the scan is essentially S-shaped and symmetric). The imaginary part of the nonlinear phase shift ∆φimag. can be obtained either from the asymmetry of the closed-aperture scan (with ∆φimag. = ∆φreal /4π) or from the depth of a dip in the open-aperture scan that is directly related to the value of ∆φimag. . Figure 4.7 shows examples of closed- and open-aperture scans for a germanium arsenic selenide film. The relation between the nonlinear phase shift and the nonlinear refractive index can be written as (4.1) ∆φ = 2πn2 ILeff /λ, where I is the light intensity, Leff is the effective sample thickness (e.g., corrected for one-photon absorption, Leff = (1 − e−αL )/α, and α is the linear absorption coefficient). Knowledge of the light intensity can be used for conversion from phase-shift values to the nonlinearity values. It is, however, more convenient to perform the measurements in a relative manner. They therefore calibrated the values of the NLO parameters by performing measurements of the nonlinear phase shift for a fused silica plate for which a value n2 = 2×10−16 cm2 W−1 was assumed. The thickness of the fused silica substrate was 1 mm. The sign of n2 deduced from Z-scan measurements was positive. Their results show that the nonlinear response of Ge33 As12 Se55 is dominated by an induced absorption effect. The value of the real part of the nonlinearity is Re(n2 ) ≈ 2.2 × 10−13 cm2 W−1 and its imaginary part is characterized by a nonlinear absorption coefficient β2 = 5.6 × 10−8 cm W−1 . It should be mentioned that Z-scan measurements on As2 S3 films did not yield reliable signals, and hence Z-scan measurements

4.1 Measurements of Nonlinear Refractive Index n2

81

Normalized open and closed aperture transmission (a.u.)

1.2

1

0.8

0.6 −20

−10

0

10

20

Distance from focus (mm)

Fig. 4.7. Open (squares) and closed (circles) aperture Z-scan results obtained on a 2 µm Ge33 As12 Se55 film on a 1 mm silica substrate. The lines are results of numerical fitting. The spot size was 32 µm, Re(∆φ) = 0.36 rad, Tt = 12. The results show that the nonlinear response of GeAsSe is dominated by an induced absorption effect. By calibrating the Z-scans against silica, one can calculate the real and imaginary parts of the nonlinearity. The real part of the nonlinearity is approximately Re(n2 ) = 2.2 × 10−13 cm2 W−1 and the imaginary part of the nonlinearity is characterized by a nonlinear absorption coefficient β2 = 5.6 × 10−8 cm W−1 (after [146])

1.3 Closed Aperture

1.2

Z-scan signal

Open Aperture

1.1 1 0.9 0.8 0.7 −15

−10

−5

0

5

10

15

Position (mm) Fig. 4.8. Open- and closed-aperture Z-scan signals from a 2-mm-thick bulk sample of As2 S3 glass measured at 1,550 nm. Note the absence of any nonlinear absorption in the open-aperture scan (after [147])

82

4 Measurement of Nonlinear Optical Constants

9.010−14 8.010−14

n2(cm2/W)

7.010−14 6.010−14 5.010−14 4.010−14 3.010−14 2.010−14 1.010−14 1000

1100

1200

1300

1400

1500

1600

Wavelength (nm) Fig. 4.9. Nonlinear refractive-index versus wavelength for As2 S3 glass (after [147])

for the case of As2 S3 were performed using glass disks [147]. Figure 4.8 shows the results of one such Z-scan measurement. Zakery et al. [147] have used Z-scan measurements to find the variation of nonlinearity n2 with wavelength for As2 S3 glass. The values varied slightly with wavelength, reaching a maximum for wavelengths around 1,300 nm (200 × n2 (silica)) and decreasing to around 100 × n2 (silica) at 1,500 nm (Fig. 4.9). Figure 4.10 shows an example of a nonphase-matched DFWM signal obtained for intensities in the range of 1–60 GW cm−2 [148]. These measurements were made using a femtosecond laser with hundred-femtosecond pulses. The energy per pulse was up to 35 µJ and the spot size on the sample was approximately 200 µm. As2 S3 films up to 4 µm thick were used in these measurements. Normally two signals were monitored. One was generated as a result of phase-matched interaction of the three incident beams and the other was one of the nonphase-matched signals generated mostly by the arsenic sulfide film. The signals recorded at lower intensities show essentially only the instantaneous response similar to that obtained from the bare silica substrate. A very weak tail of the signal appears at high intensities, probably due to the formation of permanent gratings that are gradually formed in the material that distort the background signal. Figure 4.11 shows a comparison of the power dependence of the phasematched and nonphase-matched DFWM signal for a 4 µm thick film. As expected for Kerr nonlinearity, the dependences are roughly cubic. The modulus of the nonlinear refractive index of films was calculated from a comparison of the DFWM signals with those for bare silica substrates. The modulus of n2 was found from the above measurements to be 2.7×10−14 cm2 W−1 at 800 nm, which was in good agreement with literature data.

4.1 Measurements of Nonlinear Refractive Index n2

83

1

54

Intensity

30

21 7 0.1

6 1.6 1.1 0

40

80

120

Delay (femtoseconds)

Fig. 4.10. Degenerate four-wave mixed (DFWM) signal versus the delay time for an As2 S3 film. The numbers on the graphs correspond to the input pulse intensities (in GW cm−2 ) (after [148])

Fig. 4.11. Double-logarithmic plot of power dependencies of the DFWM signals (diamonds, phase matched; squares, nonphase matched) from a 4-µm thick As2 S3 film. The line shows a theoretical cubic dependence (after [147])

Bindra et al. [149] have used Z-scan measurements at 1,550 nm and obtained n2 values of 18 × 10−19 m2 W−1 , 9.2 × 10−19 m2 W−1 , and 3.8 × 10−19 m2 W−1 for As2 S3 , GeS2 , and TeO2 , respectively. Smektala et al. [150] have used Z-scan measurements at 1.06 µm and found that the nonlinear refraction properties found for As2 S3 were consistent with the literature. Moreover, a nonlinear refractive index four times that of As2 S3 has been measured for a bulk Ge10 As10 Se80 sample.

84

4 Measurement of Nonlinear Optical Constants

Nonlinear optical properties of As–S–Se chalcogenide glasses were measured [151] by the Z-scan technique at 1.6 µm, and values of n2 up to 400 times that for silica were observed. Such large values of n2 for glasses with small As/(S + Se) molar ratios are correlated with the presence of covalent, homopolar Se–Se bonds in the glass structure as identified by Raman spectroscopy, and cannot be attributed to any red shift in the absorption edge or to a resonant effect. Spalter et al. [152] have tested the ability of Ge25 Se75 chalcogenide films for ultrafast AOS and investigated nonlinear pulse propagation in their photodarkened samples. Nearly transform limited, 270-fs pulses were coupled into the carefully cleaved guides. Pulse energies ranged up to 1.04 nJ in front of the input facet at a repetition rate of 13.5 MHz. The overall input-to-output transmission was 13%. Pulse spectra were measured with an optical spectrum analyzer. The results for Ge25 Se75 indicate that, at low energies, the output spectrum is identical to the input spectrum. Increasing the energy results in significant spectral broadening, as well as in an oscillating structure, as expected from SPM. By comparing their spectrum to numerical simulations, they inferred a peak nonlinear phase shift of 3.5π, from which a value of n2 ≈ 1.5 × 10−14 cm2 W−1 (58 times the value for silica) was obtained, which agreed well with independent Z-scan measurements. The nonlinear refractive index n2 of binary La2 S3 –Ga2 S3 and ternary MS–La2 S3 –Ga2 S3 (MS=Ag2 S and Na2 S) glasses were measured at 532 nm by the Z-scan method [133]. The n2 of the glasses increased with increasing La2 S3 content in the binary glasses, or decreased with the addition of Ag2 S or Na2 S, respectively, in the ternary glasses. These results qualitatively agree with measurements by the THG method. An optical power limiter utilizing the TPA phenomena acts as a protective element to restrict the irradiance of light pulses upon sensitive optical components, or as a regulator to smooth optical transients. Because the La2 S3 –Ga2 S3 glasses possess large β values at 532 nm, they could be promising materials for use as optical power-limiting materials. Figure 4.12 shows the result of the output transmittance versus the input intensity. The open circles represent the experimental data. The broken line is the hypothetical linear relation between Iout and Iin when TPA is not present. As seen in Fig. 4.12, the experimentally transmitted intensity was suppressed below the broken line due to the power-limiting effect of TPA. The solid line in Fig. 4.12 is the theoretical prediction using β = 41.3 cm GW−1 [133]. Lenz et al. [142] have measured n2 using Z-scan experiments at 1.5 µm for a number of chalcogenide compositions. Their results indicate that values of n2 500 times that of silica for As2 S3 glasses are possible. Based on their results, and assuming a 1 pJ, 1 ps pulse and an effective mode area of 1 µm2 , resulting in a peak intensity of 100 MW cm−2 , would require a 5–8 cm long device for AOS applications. They assert that the above numbers indicate that small, integrated devices are possible and can be combined with existing silicon optical-bench technology.

4.1 Measurements of Nonlinear Refractive Index n2

85

Transmitted intensity/GW.cm−2

0.10 -10La2S3.60Ga2S3 glass 0.08 0.06 0.04 0.02 0 0

0.1

0.2

0.3

Input intensity/GW . cm−2

Fig. 4.12. Optical-limiting behavior for 40La2 S3 .60Ga2 S3 glass of thickness 0.1039 cm. The open circles represent the experimental data, the solid line is the theoretical curve and the broken line is what is expected in the absence of TPA (after [133])

Bulk chalcogenide glasses spanning a range of compositions have been measured for their nonlinear index of refraction n2 at 1.3 and 1.55 µm using the Z-scan technique [153]. The figure of merit T (T = βλ/n2 ) was calculated for each glass and used as an indication of potential of these glasses for alloptical switching applications. SPM measurements for the As40 S60 samples were performed using an additive pulse mode-locked color-center laser producing 500 fs pulses. The SPM was estimated at π/2 assuming a 500 W peak power in the waveguide, suggesting a value of n2 = 8 ± 4 × 10−15 cm2 W−1 , which is within the error of the Z-scan measurements. Smolorz et al. [154] presented measurements of third-order optical nonlinearities in heavy-metal oxide (containing Ga, La, Bi, Pb) and sulfide glasses, using Z-scan and spectrally resolved two-beam coupling (SRTBC) techniques. The nonlinear index of refraction was found to increase with the sulfide content, and the nuclear contribution to the nonlinearity was found to be approximately constant at (15 ± 3)%. The largest nonresonant nonlinear index of refraction occurs in 35La2 S3 .65Ga2 S3 , which is 30 times larger than the nonlinear index of refraction of fused silica. Petkov et al. [155] have developed a formula for predicting the nonlinear refractive index n2 for chalcogenide glasses from the dispersion of n, and which enables n2 to be related to structural parameters. Using the various formulae and the measured values of n, n2 values for these materials have been predicted. The results [156] indicate that glasses with compositions near As42 S58 or As2 S3 Tl0.13 may, after UV exposure, exhibit significantly (about 25%) larger values of n2 than as-deposited As40 S60 . Third- and second-order nonlinear optical properties of chalcogenide glasses in the (Ge–Se–S–As) system have been studied [157]. The Z-scan and

86

4 Measurement of Nonlinear Optical Constants

Mach-Zehnder interferometry measurements of the nonlinear refractive-index n2 performed at 1,064 nm are in good agreement. The evaluation of the nonlinear refractive-index values has been correlated to the structure of the glasses, more particularly in the case of binary Ge–Se glasses when n2 is a function of the length of selenium chains in the glass. Third-order nonlinearities as high as 24 × 10−18 m2 W−1 (850 times the nonlinearity of silica glass) have been obtained [157]. The third-order nonlinear susceptibility χ(3) of homogeneous glasses has been studied by Nasu et al. [158]. χ(3) roughly depends on the refractive index but is not uniquely determined. The largest χ(3) among the homogeneous glasses is 1.96 × 10−19 m2 V−2 for As–S–Se glasses, being almost 103 times larger than that of silica. Boudebs [159] has investigated both experimentally and theoretically the optical nonlinearities of two chalcogenide glasses (As2 Se3 and As2 S3 ). Experimental data given by the spatially resolved Mach-Zehnder technique clearly indicate that the samples used in their experiments cannot be described with the usual third-order nonlinear theory. Consequently, they have constructed a model based on the existence of both a cubic and quintic nonlinear index. The evolution of the resulting nonlinear index coefficient as a function of the intensity is in good agreement with the experimental data. A fit allowed the determination of values of the nonlinear index coefficients. In particular, a negative quintic nonlinear index coefficient has been demonstrated. Rangel Rojo et al. [160] report a study of the third-order optical nonlinearities of amorphous selenium using picosecond pulses at 1.064 µm, from a mode-locked Nd:YAG laser. The Z-scan technique was used to resolve the absorptive and refractive contributions to the nonlinear response of the material, including their sign. The chosen wavelength lies to the lower photon energy side of the absorption edge for the material studied; the interaction is therefore nonresonant and electronic in origin. They measured n2 = −0.06 cm2 GW−1 (3 orders of magnitude larger than for As2 S3 and Ge33 As12 Se55 ), with negligible TPA. Kim et al. [161] deposited thin films of As2 Se3 glass by thermal evaporation. These films were found to have significantly large third-order nonlinear coefficients at the wavelength of 633 nm. The self-focusing effect results in a reduced beam size, as confirmed by a laser beam profiler. This indicates that thin-film As2 Se3 is a strong candidate for reducing the laser-beam spot size and thus can enhance the density of phase-change optical disks. Wang et al. [156] report measurements of third-order optical nonlinearity of 90GeS2 –5GeS3 –5CdS (in mole%) chalcogenide glass using the femtosecond time-resolved optical Kerr-gate technique at 820 nm. The third-order nonlinear susceptibility was estimated to be as large as 1.4 × 10−20 m2 V−2 . The full width at half maximum of the Kerr signal was 150 fs, implying that the sample had a response faster than 120 fs. Its response was dominantly assigned to the ultra-fast distortion of the electron cloud.

4.1 Measurements of Nonlinear Refractive Index n2

87

Munzar et al. [162] have prepared amorphous Gex S1−x films by thermal evaporation. Using Miller’s generalized rule (χ(3) (ω4 , ω3 , ω2 , ω1 ) = Aχ(1) (ω4 )χ(1) (ω3 )χ(1) (ω2 )χ(1) (ω1 ), where ω4 = ω1 + ω2 + ω3 and A is a quantity that is assumed to be frequency independent and nearly the same for all materials [163]), values of the third-order nonlinear susceptibility were estimated. They report a value of 6.72 × 10−21 m2 V−2 for the composition x = 0.36. Kobayashi et al. [164] have carried out THG measurements and optical Kerr-shutter operation using As2 S3 glass. The value of χ(3) for the As2 S3 glass obtained by THG at the fundamental wavelength of 2.1 µm was 1.4 × 10−19 m2 V−2 (300 times that for silica glass). THG measurements also showed the spectra of χ(3) , which becomes larger at shorter wavelengths where the TH wave is absorbed, which indicates the three-photon resonance effect. Operation of an OKS using this glass has achieved a large value of n2 = 6.8 × 10−18 m2 W−1 . The dependence of the output power on the gate light power has also been measured. The efficiency is affected by the gate-light absorption, the TPA, and the loss, which is thought to be due to excited carriers. The values of n2 obtained from the Kerr-shutter operation agree with the values estimated from THG measurements. Petkov et al. [165] have used a semi-classical model of the simple harmonic oscillator given by Boling et al. [166]. They derived a relation between the nonlinear refractive index n2 , the resonance frequency, ω0 , and the product N S (N is the density of polarizable ions and S is the oscillator strength). They showed that ω0 and N S could be related to the parameter nd , and the Abbe number νd = (nd − 1)/(nf − nc ), where nd , nf , and nc are the linear refractive indices at the following standard wavelengths λf = 486.13 nm and λd = 587.56 nm, and λc = 656.27 nm. Boling et al. [166] derived two formulae for predicting n2 to an accuracy of ∼20% or better using the values of the linear refractive index, n. However, it is difficult to find νd and nd values for IR-transmitting glasses, such as chalcogenide glasses, since they are relatively absorbing in the visible. A better fit is obtained by fitting the complete dispersion curves of these materials using the Wemple–Di Domenico model [167] and then Boling’s formula. Using the parameters E0 , the oscillator energy, and Ed , the dispersion energy, the following expression for n2 could be used: √ n2 = 3gS(n2 + 2)1.5 (n2 − 1)2 ¯h2 e2 /12 nm Ed E02 , (4.2) where n is the linear refractive index at long wavelengths, S the oscillator strength, h ¯ is Planck’s constant divided by 2π, g is the anharmonicity parameter, e and m are the electron charge and mass, respectively. Values of n2 before and after exposure of thin As–S and As–S–Tl films to UV illumination were compared. It was found that n2 increases with increasing As content in the films (unexposed and exposed), passing through a maximum for the composition As42 S58 [165]. This correlates with the data obtained for nc . The nonlinear index for unexposed thin As–S–Tl films increases with increasing Tl content. n2 also increases after illumination.

88

4 Measurement of Nonlinear Optical Constants

Ganeev et al. [168] presented the characterization of nonlinear optical parameters of chalcogenide films {As2 S3 , As20 S80 , 2As2 S3 /As2 Se3 , 3As2 S3 / As2 Se3 }. 2As2 S3 /As2 Se3 represents a multilayered film structure in which the thickness of the As2 S3 layer is twice that of the As2 Se3 layer. They measured nonlinear refractive indices and TPA coefficients using the Z-scan method at the wavelength of an Nd:YAG laser at 1,064 nm and its second harmonic at 532 nm. Ganeev et al. [169] have investigated the properties of As2 S3 and CdS nanoparticle aqueous solutions prepared by laser ablation. Nonlinear optical characteristics of these solutions were studied by the Z-scan technique using an Nd:YAG laser and its second harmonic (λ = 532 nm, tp = 55 ps). Nonlinear refractive indices, nonlinear absorption coefficients, and third-order nonlinear susceptibilities of these solutions were measured. It was shown that nonlinear refractive indices of As2 S3 and CdS nanoparticles decreased with a growth in laser intensity. High nonlinear optical susceptibilities of such structures were attributed to size-related effects. A value of n2 for CdS solution was calculated to be ∼0.86 × 10−19 m2 W−1 at an intensity of 3.9 × 109 W cm−2 . For As2 S3 solution, a value of n2 = 1.46 × 10−18 m2 W−1 was calculated at an intensity of 2.94 × 109 W cm−2 . Tichy et al. [170] have used parameters of the Wemple–DiDomenico singleoscillator model for the linear refractive-index dispersion and Miller’s rule and estimated the values of the third-order nonlinear optical susceptibility for several amorphous chalcogenide thin films. Estimated values are large (up to χ(3) = 1.05 × 10−19 m2 V−2 for Ge2 As40 Se58 thin films) and are comparable with χ(3) values observed for other chalcogenide systems [139]. Cherukulappurath et al. [171] have investigated nonlinear coefficients of chalcogenide glasses containing different amount of tellurium (Ge10 As10 Se80−x Tex , x = 0, 10, 15, 20). A comparison of the measurements of the nonlinear coefficient obtained by Z-scan with values given by other methods show that the agreement is good at the same incident intensity. It was seen that the nonlinear refraction coefficients measured were among the largest values reported for chalcogenide glasses (n2 ∼ 20 × 10−18 m2 W−1 ). Furthermore, they found that the addition of tellurium does not enhance significantly the nonlinear refraction coefficient. Ogusu et al. [172] have prepared Agx (As0.4 Se0.6 )100−x glasses and measured their nonlinear optical properties at 1.05 µm (see Fig. 4.13). The measured nonlinear refractive index of the glass with x = 20 at.% is approximately 2–4 times as large as that of As2 Se3 glass and ranges from 2,000 to 27,000 times as large as that of fused silica, depending on the incident intensity. Although the figure of merit F (F = 2λβ/n2 ) of the samples tested at 1.05 µm does not satisfy the standard criterion (F 0), the GVD and SPM effects can be used for pulse compression. It is sometimes necessary to include the third-order term proportional to β3 in (5.41). This happens if the pulse wavelength nearly coincides with the zero-dispersion wavelength λD ,β2 ≈ 0; the β3 term then provides the dominant contribution to the GVD effects [259]. Also, for ultrashort pulses

122

5 Optical Nonlinearities in Chalcogenide Fibres

(width T0 < 1 ps), it is necessary to include the β3 term even when β2 = 0. In a fiber-optic communication system, information is transmitted along a fiber using a coded sequence of optical pulses whose width is determined by the bit rate, B, of the system. Dispersion-induced broadening of pulses is undesirable because it interferes with the detection process and leads to errors if the pulse spreads outside its allocated bit slot (TB = 1/B). Clearly, GVD limits the bit rate for a fixed transmission distance L [260]. There are several schemes to compensate for GVD dispersion. First, modern fiber-optic communication systems operating near 1.55 µm reduce the GVD effects using dispersion-shifted fibers designed such that the minimum-loss wavelength and the zero-dispersion wavelength nearly coincide. Although operation at the zero-dispersion wavelength is most desirable, other considerations may preclude such a design, because at most one channel can be located at the zero-dispersion wavelength in a wavelength-divisionmultiplexed (WDM) system. The technique of dispersion management provides a solution to this problem. It consists of combining fibers with different characteristics such that the average GVD of the entire fiber link is quite low. When the bit rate of a single channel exceeds 100 Gb s−1 , ultrashort pulses (width ∼ 1 ps ) would be used in each bit slot. In this case, the pulse spectrum becomes broad enough that it is difficult to compensate GVD over the entire bandwidth of the pulse. The solution to this problem is provided by fibers, or other devices, designed such that both β2 and β3 are compensated simultaneously [261–275]. Fiber gratings, liquid–crystal modulators, and other devices can also be used for this purpose [267–273]. When both β2 and β3 are nearly compensated, propagation of femtosecond optical pulses is limited by the fourth-order dispersion effect governed by the parameter β4 . Compensation of dispersion up to fourth-order has been achieved [275].

5.5 Applications Various kinds of third-order nonlinear optical materials have been studied for their suitability of use in all-optical switching [276], and pulse compression [277]. Among them, glass fibers offer the advantage of reduced operating power because of the long interaction length. Silica fibers have been extensively studied because of their low losses [278]. However, their nonlinearity is weak and they require very long fibers (>100 m). So other highly nonlinear glasses are needed. Third-harmonic generation (THG) measurements have revealed that As2 S3 glass has a χ(3) two orders of magnitude higher than that of silica glass [279]. A very efficient optical Kerr effect in As2 S3 -based chalcogenide glass fiber has been observed [280] and efficient all-optical switching using a small-core fiber only a few meters long has been demonstrated [281]. For practical applications, some other nonlinear properties of As2 S3 glass, such as the response time, are needed. It should be noted that enhancement of third-order nonlinear effects other than the optical

5.5 Applications

123

Kerr effect, such as stimulated Raman scattering (SRS) and two-photon absorption (TPA), would be predicted. In silica fibers, it is well known that SRS can lead to delayed nonlinear response [258, 282]. Also, TPA limits the transmittable gate power and thereby limits the obtainable phase shift [283, 284]. The third-order nonlinear properties of As2 S3 -based glass fibers have been studied at a wavelength of around 1.55 µm [285]. An ultrafast response less than 100 fs was obtained by time-resolved pump-probe measurements and spectrum broadening due to SPM. It was claimed that SRS was observed in a fiber for the first time. The Stokes shift was in good agreement with the spontaneous Raman data and the Raman-gain coefficient was 4.4 × 10−12 m W−1 , which is two orders of magnitude higher than that of silica fiber. As expected, the Raman gain led to a retarded nonlinear response, with a relaxation time of about 97 fs. A weak transmission change due to TPA was observed and the TPA coefficient was estimated to be 6.2 × 10−15 m W−1 . It was asserted that, from the results obtained, As2 S3 -based glass fibers offer an ultrafast response and low TPA characteristics, and are therefore appropriate for nonlinear optical media at communication wavelengths. All-optical switching using As2 S3 -based fiber was demonstrated in the Kerr-shutter and in a nonlinear optical loop mirror (NOLM) configuration. In the first demonstration of the Kerr shutter using a 50-cm-long fiber at 1.3 µm, the switching power was 14 W. Afterward, the switching power was reduced to 3 W in a 1-m-long small-core fiber, using laser diode driving sources and assistance of an erbium-doped fiber amplifier (EDFA) at 1.55 µm. The switching power was reduced to 0.4 W using a 4-m-long low-loss fiber in NOLM [256]. One of the simplest ways to reduce further the switching power is to increase the interaction length by using a longer fiber. However, the large GVD of an As2 S3 fiber causes walk-off between the gate and the signal, which limits the effective interaction length. To overcome this limitation of the switching power or switching speed due to GVD, some kind of compensation technique is needed. Chirped gratings in particular can be used for dispersion compensation [286]. Since glasses lack a center of inversion symmetry, and thus have no secondorder nonlinear susceptibility, they should not show SHG. However, undoped and Pr-doped GaLaS glasses have exhibited SHG [287]. This SHG may be due to crystallization or the effect of frozen-in electric fields. Electric poling has been successfully used to obtain SHG in silica-based fiber systems [288]. It is not unreasonable to expect similar results in chalcogenide fibers except that, since the electrical conductivity is lower for chalcogenide glasses, the (breakdown) electric field will be lower. Dixit et al. [289] have investigated a Fabry–Perot (FP) filter and a nonlinear fiber-loop mirror (NFLM) for their suitability as photonic switching devices. They have found from their analysis that fibers made of doped chalcogenide glasses hold promise for applications. The advantage is seen to be more in the case of FP filters than in NFLM. It is seen that Se-based chalcogenide

124

5 Optical Nonlinearities in Chalcogenide Fibres

glass fibers are a better choice in Fabry–Perot switching devices even with attenuation levels of α = 170 dB m−1 . This is because the switching power is 7.5 W for a length of 2 mm, in contrast to 23 kW for conventional silica glass with a length of 2 cm. If α is reduced to values such as 3 dB m−1 , as in As2 S3 -based glass fibers, the switching power can be brought down to 442 W, with a device length of 2 cm. These authors state that, in contrast, chalcogenide glass fiber may not be a better option for NFLM unless the present loss values of the order of 3 dB cm−1 in this material are reduced to a greater extent. The present values of loss coefficients in these glasses necessitate the use of very small loop lengths (∼15 cm). If the loss can be decreased to a value of 0.2 dB m−1 , then the improved figure of merit would bring down the switching power with increased lengths (∼2 m), thus adding to its advantages. If fluctuations in composition and density, which normally lead to large attenuation levels, can be controlled, thus leading to materials with higher values of nonlinear coefficient and lower values of α, the overall figure of merit can be improved. In recent years, great interest has been generated by the development of holey or microstructured optical fibers [290]. The large and controllable variations of transverse refractive index offered by these fibers provide new opportunities for the control and guidance of light [291, 292]. To date, holey fibers are usually made from conventional silica glasses. The significant advantage for novel glasses is that different core and cladding glass compositions are no longer required. This relaxes the fabrication difficulties associated with core and cladding glasses with differing thermal properties. It has been difficult in the past to fabricate low-loss single-mode compound glass fibers due to problems arising from the different physical properties of the core and cladding materials. The varied heating steps required to fabricate single-mode compound fiber leads to the promotion of crystallization, and hence inducing of losses. Holey fibers can be made using a single heating step, reducing crystallization problems, and significantly reducing the fiber loss. Hence, holey fibers provide a new route toward the successful development of low-loss single-mode compound glass optical fibers. Planar channel waveguide devices based on the GLS glass system have the potential for use in glass waveguide lasers, wavelength multiplexers, and optical switching/splitting in the IR [290]. Further potential for optical couplers (such as directional couplers) in the IR would allow the important applications of power division and wavelength demultiplexing to be realized. Further devices can be derived from the higher index contrast possible with GLS holey fibers allowing for fibers with very high numerical aperture NA (well in excess of unity). As a result of improvement in pump confinement, tight focusing and shorter devices and lower thresholds are possible. Third-order Kerr nonlinearities and Raman gain have been studied experimentally in high-purity As2 Se3 optical fibers for wavelengths near 1.55 µm [293]. Raman-gain measurements on the same fiber used for the n2 measurements were carried out with a CW pump at a wavelength of 1,540 nm. Raman

5.5 Applications

125

1.3

Pout/Pin

1.2

1.1

1.0 1580

1590

1600

1610

1620

Wavelength (nm)

Fig. 5.7. Raman-gain spectra measured in an 85-cm-long As–Se fiber (solid curve) compared with the Raman spectrum for the bulk material (dotted curve). The ratio of the power output to the power input, normalized to the ratio well removed from the Raman-gain region, is shown as a function of wavelength (after [293])

gain was expected at the Raman frequency shift of 830 cm−1 , corresponding to a wavelength shift of 65 nm from the pump and a Raman-gain peak near 1,605 nm. The ratio of the output to input signal levels normalized to the ratio at wavelengths outside the Raman-gain spectral region is shown in Fig. 5.7 for an average pump power in the fiber of 190 mW. We define the Raman gain by: (5.51) Iout = Iin exp(gR Ip L), where gR is the Raman gain in meters per watt, Ip is the pump intensity in watts per square meter, L is the fiber length in meters, and Iin and Iout are the input and output signal intensities, respectively. A value of gR ≈ 5.1 × 10−11 m W−1 , which is nearly 780 times the value for silica, is obtained. The linear Raman spectrum measured for bulk As2 Se3 is shown in Fig. 5.7 as the dotted curve. Since the gain is small (